monte carlo treatment planning - radiation...

Download Monte Carlo Treatment Planning - Radiation Dosimetryradiationdosimetry.org/files/...16-monte-carlo-treatment-planning.pdf · Monte Carlo Treatment Planning An Introduction NEDERLANDSE

If you can't read please download the document

Upload: hoanghanh

Post on 06-Feb-2018

231 views

Category:

Documents


1 download

TRANSCRIPT

  • Monte Carlo Treatment Planning

    An Introduction

    NEDERLANDSE COMMISSIE VOOR STRALINGSDOSIMETRIE

    Report 16 of the Netherlands Commission on Radiation Dosimetry

    Netherlands Commission on Radiation Dosimetry

    Subcommission Monte Carlo Treatment Planning

    June 2006

  • Monte Carlo Treatment Planning

    An Introduction

    NEDERLANDSE COMMISSIE VOOR STRALINGSDOSIMETRIE

    Report 16 of the Netherlands Commission on Radiation Dosimetry

    Authors:

    N. Reynaert

    S. van der Marck

    D. Schaart

    W. van der Zee

    M. Tomsej

    C. van Vliet- Vroegindeweij

    J. Jansen

    M. Coghe

    C. De Wagter

    B. Heijmen

    Netherlands Commission on Radiation Dosimetry

    Subcommission Monte Carlo Treatment Planning

    June 2006

  • i

    Preface

    The Nederlandse Commissie voor Stralingsdosimetrie (NCS, Netherlands Commission on

    Radiation Dosimetry) was officially established on 3 September 1982 with the aim of

    promoting the appropriate use of dosimetry of ionizing radiation both for scientific research

    and practical applications. The NCS is chaired by a board of scientists, installed upon the

    suggestion of the supporting societies, including the Nederlandse Vereniging voor

    Radiotherapie en Oncologie (Netherlands Society for Radiotherapy and Oncology), the

    Nederlandse Vereniging voor Nucleaire Geneeskunde (Netherlands Society for Nuclear

    Medicine), the Nederlandse Vereniging voor Klinische Fysica (Netherlands Society for

    Clinical Physics), the Nederlandse Vereniging voor Radiobiologie (Netherlands Society for

    Radiobiology), the Nederlandse Vereniging voor Stralingshygine (Netherlands Society for

    Radiological Protection), the Nederlandse Vereniging voor Medische Beeldvorming en

    Radiotherapy (Netherlands Society for Medical Imaging and Radiotherapy), the Nederlandse

    Vereniging voor Radiologie (Netherlands Society for Radiology) and the Belgische

    Vereniging voor Ziekenhuisfysici/Socit Belge des Physiciens des Hpitaux (Belgian

    Hospital Physicists Association.

    To pursue its aims, the NCS accomplishes the following tasks: participation in dosimetry

    standardisation and promotion of dosimetry intercomparisons, drafting of dosimetry

    protocols, collection and evaluation of physical data related to dosimetry. Furthermore the

    commission shall maintain or establish links with national and international organisations

    concerned with ionizing radiation and promulgate information on new developments in the

    field of radiation dosimetry.

    Current members of the board of the NCS:

    S. Vynckier, chairman

    B.J.M. Heijmen, vice-chairman

    E. van Dijk, secretary

    J. Zoetelief, treasurer

    A.J.J. Bos

    A.A. Lammertsma

    J.M.Schut

    F.W. Wittkmper

    D. Zweers

  • ii

    Monte Carlo Treatment Planning: An Introduction

    This report was prepared by a subcommittee of the Netherlands Commission on Radiation

    Dosimetry (NCS), consisting of Belgian and Dutch scientists.

    Members of the subcommittee:

    N. Reynaert, chairman

    S. van der Marck

    D. Schaart

    W. van der Zee

    M. Tomsej

    C. Van Vliet-Vroegindeweij

    J. Jansen

    M. Coghe

    C. De Wagter

    B. Heijmen

    Monte Carlo Treatment Planning: An Introduction

    Report 16 of the Netherlands Commission on Radiation Dosimetry (NCS)

    June 2006

    NCS, Delft, The Netherlands

    ISBN 90-78522-01-1

    For more information on this and other NCS Reports, see http://www.ncs-dos.org

  • iii

    User guide

    This report presents an overview of the literature for physicists in radiotherapy departments

    who intend to buy/use/customise a Monte Carlo treatment planning system for electron

    and/or photon therapy. The report focuses on commissioning, selection of treatments

    requiring Monte Carlo, variance reduction techniques, accelerator head modelling, patient

    modelling (conversion of CT Hounsfield units), hardware requirements and the required

    knowledge to operate an MCTP system. In addition an overview of existing Monte Carlo dose

    engines and MCTP systems is given.

    The report consists of three main parts.

    The first part provides insight in the Monte Carlo method for dose calculations. An overview

    of general purpose Monte Carlo codes, used in the field of electron and photon dosimetry, is

    given. An extensive description of modelling of electron and photon transport and the usage

    of cross sections is presented.

    The second part deals with MCTP specific topics such as CT conversion, linac head

    modelling, scoring, variance reduction, Monte Carlo based treatment planning (optimisation),

    and 4D planning.

    The third and final part focuses on practical aspects. It provides an overview of Monte Carlo

    dose engines used for Monte Carlo treatment planning, an overview of commercial MCTP

    systems, and guidelines on benchmarking of these systems (focussing on MC specific

    benchmarks).

  • iv

    Contents PREFACEI

    USER GUIDE....III

    CONTENTS..IV

    SUMMARY ........................................................................................................................1

    ABBREVIATIONS ..................................................................................................................3

    1 INTRODUCTION.....................................................................................................5

    PART I: INTRODUCTION TO MONTE CARLO......................................................................6

    2 MONTE CARLO FOR SOLVING NUMERICAL PROBLEMS ..................................7

    2.1 COMPARISON WITH ANALYTICAL AND NUMERICAL APPROACHES ...............7

    2.2 MONTE CARLO DOSE CALCULATIONS...............................................................7

    2.3 EXAMPLE: AN 8 MEV ELECTRON HITTING THE LINAC TARGET.......................8

    3 BASIC ELEMENTS OF A MONTE CARLO CODE FOR DOSE CALCULATIONS 11

    3.1 PHYSICS MODELS ..............................................................................................11

    3.2 INTERACTION DATA TABLES.............................................................................11

    3.3 RANDOM NUMBER GENERATOR ......................................................................12

    3.4 GEOMETRY .........................................................................................................12

    3.5 MATERIAL COMPOSITION..................................................................................12

    3.6 SOURCE DEFINITION..........................................................................................12

    3.7 SCORING .............................................................................................................13

    3.8 VARIANCE REDUCTION AND APPROXIMATIONS ............................................13

    4 A BRIEF HISTORY ...............................................................................................14

    4.1 GENERAL PURPOSE CODES.............................................................................15

    5 GENERAL PURPOSE MONTE CARLO CODES IN RADIOTHERAPY.................17

    5.1 EGS ......................................................................................................................17

    5.2 MCNP ...................................................................................................................19

    5.3 PENELOPE...........................................................................................................20

    5.4 GEANT .................................................................................................................21

    6 RATIONALE FOR MONTE CARLO TREATMENT PLANNING.............................23

    6.1 REQUIREMENTS ON UNCERTAINTY IN TREATMENT PLANNING...................23

    6.2 WHY MONTE CARLO TREATMENT PLANNING.................................................24

    6.3 PHANTOM EXPERIMENTS..................................................................................25

    6.4 COMPARISONS FOR CLINICAL CASES.............................................................28

    6.5 CONCLUSIONS....................................................................................................38

  • v

    PART II: FUNDAMENTALS OF MONTE CARLO ................................................................40

    7 MODELLING OF PARTICLE TRANSPORT..........................................................41

    7.1 PHOTON TRANSPORT........................................................................................41

    7.2 ELECTRON TRANSPORT....................................................................................43

    7.3 INTERACTION DATA TABLES.............................................................................48

    8 GEOMETRY AND MATERIAL SPECIFICATION ..................................................59

    8.1 VOLUMES ............................................................................................................59

    8.2 VOXELISED PHANTOMS.....................................................................................59

    8.3 CONVERSION OF CT NUMBERS INTO TISSUE PARAMETERS .......................59

    9 ACCELERATOR MODELLING .............................................................................66

    9.1 GENERAL ASPECTS ...........................................................................................66

    9.2 MODELLING OF THE LINAC HEAD.....................................................................67

    9.3 VIRTUAL SOURCE MODEL .................................................................................68

    9.4 BEAM MODIFIERS...............................................................................................71

    10 DOSE SCORING ..................................................................................................75

    10.1 DOSE DETERMINATION .....................................................................................75

    10.2 SCORING GRIDS.................................................................................................76

    10.3 SPATIAL RESOLUTION .......................................................................................77

    10.4 CONVERSION OF MONTE CARLO RESULTS TO DOSE TO WATER ...............78

    11 VARIANCE REDUCTION TECHNIQUES AND APPROXIMATIONS ....................80

    11.1 INTRODUCTION...................................................................................................80

    11.2 VARIANCE OF A MONTE CARLO CALCULATION..............................................81

    11.3 VARIANCE REDUCTION TECHNIQUES .............................................................81

    11.4 RISKS OF VARIANCE REDUCTION ....................................................................89

    11.5 DENOISING..........................................................................................................91

    12 MONTE CARLO TREATMENT PLANNING ..........................................................96

    13 4D MONTE CARLO DOSE CALCULATIONS .....................................................101

    PART III: MONTE CARLO TREATMENT PLANNING IN PRACTICE.................................106

    14 MONTE CARLO DOSE CALCULATION ENGINES FOR TREATMENT

    PLANNING..............................................................................................................107

    14.1 PIONEERING WORK .........................................................................................107

    14.2 DPM....................................................................................................................109

    14.3 MCDOSE/ MCSIM ..............................................................................................110

    14.4 VMC, XVMC, VMC++..........................................................................................110

  • vi

    14.5 PEREGRINE.......................................................................................................112

    14.6 MACRO MONTE CARLO (MMC)........................................................................113

    14.7 DOSE ENGINES SERVING AS COMMISSIONING TOOL .................................114

    15 AVAILABLE COMMERCIAL MCTP SYSTEMS...................................................116

    16 MONTE CARLO SPECIFIC ISSUES OF COMMISSIONING ..............................118

    16.1 INTRODUCTION.................................................................................................118

    16.2 PARTICLE SOURCE AND BEAM MODIFIERS ..................................................119

    16.3 SEGMENTATION ...............................................................................................120

    16.4 NORMALIZATION / MU DETERMINATION........................................................120

    16.5 VARIANCE REDUCTION....................................................................................121

    16.6 LITERATURE DATA ON MCTP VERIFICATION ................................................121

    16.7 CONCLUSION....................................................................................................126

    17 RECOMMENDATIONS.......................................................................................127

    17.1 COMPARISON OF DIFFERENT DOSE ENGINES.............................................127

    17.2 COMMISSIONING ..............................................................................................128

    17.3 CT CONVERSION ..............................................................................................129

    17.4 CONVERSION OF DOSE TO MEDIUM TO DOSE TO WATER .........................129

    17.5 VARIANCE REDUCTION TECHNIQUES AND APPROXIMATIONS ..................129

    17.6 DENOISING........................................................................................................130

    18 CONCLUSION....................................................................................................131

    REFERENCES...................................................................................................................133

    APPENDICES ....................................................................................................................159

    APPENDIX A. AN EXAMPLE TO ILLUSTRATE DIFFERENCES BETWEEN THE

    MONTE CARLO TECHNIQUE AND ANALYTICAL AND NUMERICAL APPROACHES.....160

    A.1 ANALYTICAL TECHNIQUE ................................................................................160

    A.2 NUMERICAL TECHNIQUE .................................................................................161

    A.3 MONTE CARLO TECHNIQUE............................................................................162

    A.4 SUMMARY..........................................................................................................164

    APPENDIX B: RANDOM NUMBERS IN MONTE CARLO ..................................................165

    B.1 RANDOM NUMBERS IN COMPUTERS .............................................................165

    B.2 RANDOM NUMBER GENERATORS..................................................................166

  • 1

    Summary

    The accuracy of dose calculation engines used for treatment planning in radiotherapy

    has increased steadily, ranging from calculations based on measurements, to pencil

    beam algorithms and superposition/convolution algorithms. Currently, Monte Carlo

    dose calculation engines are implemented in commercial treatment planning software

    as it is believed that the Monte Carlo method can provide an accuracy within 2-3 %. It

    is important that clinical physicists have insight in these systems, when introducing

    them into the clinic. This report tackles this acute problem by providing extensive

    information on:

    general purpose Monte Carlo codes for photon and electron dosimetry

    applications

    modelling of particle transport

    cross sections

    MCTP (Monte Carlo Treatment Planning) specific issues such as linac

    modelling, CT conversion, variance reduction techniques, scoring grids

    Recent developments such as 4D applications and MCTP optimisation

    An important question is whether the added value of MCTP is clinically relevant.

    To answer this question an extensive overview of the literature is provided. The main

    conclusion is that the MC method has important added value when compared to

    pencil beam algorithms. More information is needed when comparing MC to

    superposition/convolution algorithms, although the first experiments (comparing

    accurate Monte Carlo dose calculation engines to superposition/convolution

    algorithms) demonstrate that the MC method will become very important in clinical

    treatment planning.

    As the Monte Carlo method is, by its nature, very time consuming, a number of

    approximations have been included in commercial Monte Carlo dose calculation

    engines for treatment planning. This leads to a reduction in calculation time of several

  • 2

    orders of magnitude. The impact on the dosimetrical accuracy however is not well

    known yet. This report provides an overview of existing Monte Carlo dose calculation

    engines, focussing on applied approximations. An overview of commercial MCTP

    systems that are already available or are currently being developed is given. As

    benchmarking remains as important as for any other treatment planning system, a

    paragraph is devoted to quality control. Commercial MCTP systems can be

    benchmarked by measurements but also by comparison with accurate Monte Carlo

    dose calculation engines containing only a few approximations.

  • 3

    Abbreviations

    3D Three-Dimensional 4D Four-Dimensional AAPM American Association of Physicists in Medicine ASCII American Standard Code for Information Interchange BEAM an EGS4/PRESTA or EGSnrc/PRESTAII Monte Carlo user code CERN European Organization for Nuclear Research CSDA Continuous Slowing Down Approximation CPU Central Processing Unit CT Computed Tomography CTV Clinical Target Volume DOSXYZ an EGS4/PRESTA Monte Carlo user code DPM Dose Planning Method (MC algorithm for photons and electrons) DVH Dose-Volume Histogram EGS Electron Gamma Shower (a Monte Carlo code) ENIAC Electronic Numerical Integrator And Computer EPID Electronic Portal Imaging Device EPL Equivalent Path Length ESTRO European Society for Therapeutic Radiology and Oncology ETRAN Electron TRANsport (a Monte Carlo code) FORTRAN FORmula TRANslation (programming language) FWHM Full Width at Half Maximum GEANT GEometry ANd Tracking (a Monte Carlo code) ICRU International Commission on Radiation Units and Measurements IMRT Intensity-Modulated Radiation Therapy ITS Integrated Tiger Series (a Monte Carlo code package) KEK National Laboratory for High Energy Physics (Japan) LANL Los Alamos National Laboratory MC Monte Carlo MCDOSE an EGS4/PRESTA Monte Carlo user code MCNP3 Monte Carlo Neutron Photon (a Monte Carlo code) MCNP4 Monte Carlo N-Particle (a Monte Carlo code) MCTP Monte Carlo Treatment Planning MLC Multi-Leaf Collimator MMC Macro Monte Carlo (MC algorithm for electrons) MORTRAN Fortran pre-processor (used for EGS) MRI Magnetic Resonance Imaging MU Monitor Unit NIST National Institute of Standards and Technology NCS Netherlands Commission on Radiation Dosimetry NRC National Research Council of Canada NTCP Normal Tissue Complication Probability PB Pencil Beam PC Personal Computer PENELOPE PENetration and Energy LOss of Positron and Electrons (MC code)

  • 4

    PET Positron Emission Tomography PRESTA Parameter Reduced Electron Stepping Algorithm PTV Planning Target Volume RBE RadioBiological Effectiveness QA Quality Assurance SLAC Stanford Linear Accelerator Center SPECT Single Photon Emission Computed Tomography TPS Treatment Planning System TRUS TransRectal UltraSound TCP Tumor Control Probability VISED Visual Editor (graphical interface for MCNP) VMC Voxel Monte Carlo (MC algorithm for electrons) VMC++ MC algorithm based on VMC and XVMC XVMC MC algorithm for photons based on VMC

  • 5

    1 Introduction

    In the past decades, the sophistication of dose calculation models implemented

    in clinical radiotherapy treatment planning systems has gradually improved, together

    with available computing power in hospitals. This evolution, going from rather simple

    scatter- and inhomogeneity corrections to pencil beams and

    superposition/convolution models has resulted in continuous improvements in the

    accuracy of predicted patient doses. In superposition/convolution models, pre-

    determined Monte Carlo results are used. Full Monte Carlo dose calculations would

    therefore seem the next logical step.

    For many years it has been realised that full Monte Carlo simulations of the

    radiotherapy dose delivery process should further improve calculation accuracy. Due

    to limitations in computing power, however, this was never a realistic option in a

    clinical setting. Recently, vendors of clinical treatment planning systems have

    nevertheless started to offer Monte Carlo dose calculations. However, available

    computing power may still not allow for full Monte Carlo simulations in clinical

    practice. Approximations and simplifications to speed up the calculations may

    therefore be necessary, possibly (partially) jeopardising the advantages of full Monte

    Carlo dose calculations.

    The aim of this NCS report is to provide potential users of a clinical treatment

    planning system with an introduction in the Monte Carlo technique. Apart from

    providing an explanation of fundamental and practical aspects specific to Monte

    Carlo treatment planning, recommendations (although limited) for potential users and

    vendors are included. This report only covers external photon and electron beam

    therapy using conventional linear accelerators. Brachytherapy, hadron therapy,

    tomotherapy, robotic radiotherapy, etc., are beyond the scope of this report.

  • 6

    Part I: Introduction to Monte Carlo

  • 7

    2 Monte Carlo for solving numerical problems

    2.1 Comparison with analytical and numerical approaches

    The main difference between the Monte Carlo technique on one hand and

    analytical and numerical approaches on the other is the use of a random number

    generator and a set of probability distributions to sample parameter values for

    calculating a possible solution to the problem for a single case or event. By

    simulating many cases or events, reliable average values can be obtained. Since

    the result is an average, it is associated with a standard deviation that expresses the

    uncertainty due to the fact that the simulated number of events is less than infinite.

    This source of uncertainty is not present when analytical methods are used. Of

    course, the answer obtained with analytical methods is still associated with an

    uncertainty, arising from the common sources such as uncertainties in the input

    parameters and possible systematic errors in the model. A possible disadvantage of

    analytical methods is that solutions may be difficult to obtain for complex problems.

    (Minor) changes in the relationship between parameters, or the introduction of a new

    parameter, may create a major problem in finding a new analytical solution.

    Numerical methods are generally less sensitive to such changes. If, for

    instance, a relationship changes, the numerical algorithm can stay the same,

    because it only uses the values of the function at certain points. In Appendix A, the

    example of calculating the area of a circle with radius 1 is used to demonstrate some

    differences between the different techniques.

    2.2 Monte Carlo dose calculations

    In a Monte Carlo dose calculation, the track of each individual ionizing particle

    (in radiotherapy generally photons and electrons) through the volume of interest is

    simulated. Along its way, the particle may interact with the matter through which it is

    passing, e.g. through Compton scattering (for photons) or Coulomb scattering (for

    electrons). Using a random number generator and probability distributions for the

    different types of interaction, the program samples the distance l to the next

    interaction for a particle at a given position and with velocity vector v in a certain

  • 8

    direction. The particle is then propagated with velocity v over the distance l to the

    interaction location. Next, the program chooses the type of interaction that will take

    place. For a dose calculation, one extra step is needed. The dose is defined as the

    amount of energy deposited per unit of mass (J/kg = Gy in SI units). Therefore, for

    each interaction that is simulated, the program calculates the energy balance: the

    energy of the incoming particle(s) minus the energy of the outgoing one(s). To

    calculate the dose in a particular volume (voxel), one adds the contributions from all

    interactions taking place inside the volume, and divides this by the mass in the

    volume.

    2.3 Example: An 8 MeV electron hitting the linac target

    To illustrate some of the principles of Monte Carlo dose calculations, the

    simulation of a photon that is generated in a linac head when an 8 MeV electron hits

    the target is described. The energy distribution of the photons generated is depicted

    in the left panel of Figure 2.1 The photon energy can be determined in two ways. The

    first one is the so-called hit-or-miss method. For this method, two random numbers

    are generated, one of which, designated x, is uniformly distributed between 0.01

    and 8 (photon energy), the other, y is uniformly distributed between 0 and 1.2

    (probability density of a photon with that energy). The value of 1.2 is chosen to be

    equal to the maximum of the energy probability distribution (left panel of Figure 2.1),

    or slightly above that. The point x,y is now plotted in this probability distribution. If it

    is above the curve the target was missed, the point is rejected, and a next point is

    randomly generated. If it is below, the point is accepted, and the photon energy is x

    MeV.

  • 9

    Figure 2.1 Left panel: energy probability distribution for photons that are

    generated when an 8 MeV electron hits a linac target. Right panel: cumulative

    probability distribution generated from the left panel. The cumulative probability at a

    certain energy is the probability to generate a photon at or below that energy.

    At first glance, it may seem that there is a reasonable chance that a chosen

    point x,y will end up below the curve, yielding the hit-or-miss method rather efficient.

    However, the probability density in Figure 2.1 is plotted on a log-scale. Therefore, a

    large number of points will be rejected.

    A more efficient method for selecting photon energies is based on the

    cumulative probability distribution (right panel of Figure 2.1). For this method, values

    for the cumulative probability are randomly selected, using a single random number,

    uniformly distributed between 0 and 1. Figure 2.1 shows an example for a selected

    value of 0.732. The corresponding energy, in this case 1.3 MeV, is selected. This

    algorithm is very efficient because only one random number is needed, and each

    value results in the selection of a photon energy, i.e. there is never a miss.

    Apart from the photon energy, the angles and between the directions of the

    incoming electron and the created photon have to be selected. Also for these angles,

    probability distributions are known. Therefore, the Monte Carlo program can generate

    values for and in exactly the same way as for the energy. Once the energy and

    angles of the photon are known, the distance to the first interaction site can be

  • 10

    selected, using the attenuation coefficient ([m-1]), which is the product of the atomic

    cross sections ([m2]) of the materials that the photon encounters, and the atom

    density of these materials ([m-3]). The probability that the photon will travel a distance

    l without undergoing any interactions is then given by exp(- l ), and d l is the

    probability to interact in the interval d l . So, the probability for an interaction between

    l and l +d l is given by exp(- l )d l . Similar as for the selection of the photon

    energy (Figure 2.1), a cumulative probability curve P( l ) can now be constructed for

    selection of the (first) interaction site:

    (1.1)

    From this cumulative probability distribution of distances, the travel length l for

    a random number r in the range [0,1] can now be expressed analytically:

    (1.2)

    Here, 1-r is again a random number that is uniformly distributed between 0 and

    1; in the final step it has been replaced by a new random number, r.

    With the travel distance l to the first interaction site known, the position of the

    photon can be updated, and the type of interaction that will take place can be

    selected, based on the cross section data for the different interactions. Subsequently,

    the energies and angles of the particles that are produced in the interaction are

    generated, and the whole process is repeated until all particle energies are below a

    pre-defined cut-off energy.

    ===

    l ls edselP0

    1)( L

    )'ln(1

    )1ln(1

    1 rrler l

    ===

  • 11

    3 Basic elements of a Monte Carlo code for dose

    calculations

    3.1 Physics models

    The physics models are usually hard-coded in the Monte Carlo software.

    Photons are transported in a way that is analogue to reality. For electrons, the

    simulation of each individual interaction is very time consuming and impractical for

    radiotherapy applications. Therefore, so-called condensed history techniques have

    been introduced (section 7.2). These techniques are approximations of the real

    physics, and implementation differences exist between different codes. This may

    lead to different results, which is the main reason why these codes need to be

    thoroughly benchmarked. Even with condensed history techniques, electron transport

    often remains the most time-consuming part of radiotherapy Monte Carlo simulations.

    The user may be able to manipulate the physics modelling via a number of so-

    called transport parameters. For example, the user may enable/disable certain

    interactions and/or set the values of parameters that determine e.g. cut-off energies

    or electron step lengths. Such parameters may significantly influence a simulation.

    For example, when a particles energy decreases below the cut-off energy, it is

    discarded and the remaining energy is deposited locally. Obviously, increasing this

    parameter will increase the calculation speed, but accuracy might be lost. See

    sections 7.1 and 7.2 for details.

    3.2 Interaction data tables

    Data tables with interaction probabilities for each type of interaction for each

    element are usually provided together with a Monte Carlo program. Each of the

    Monte Carlo programs has its own format for these tables, therefore interchanging

    data tables between the various Monte Carlo programs is a non-trivial task. However,

    since these data tables are so closely linked to the Monte Carlo program, the

    installation of the program will typically also include installation of the data tables (see

    section 7.3).

  • 12

    3.3 Random number generator

    By its nature, the Monte Carlo method requires a random number generator for

    sampling the probability distributions. In computer codes, this is generally solved by

    implementing a recurrence relation. Properties such as uniformity of distribution and

    random number sequence length are crucial for the reliability of the Monte Carlo

    code. This topic is addressed in more detail in Appendix B.

    3.4 Geometry

    The geometry is to be specified by the user. Depending on the code, different

    geometric structures can be defined: planes, cylinders, spheres, cones, and

    sometimes even more complicated structures, see section 8.1. In some general

    purpose Monte Carlo codes, an (additional) scoring geometry has to be introduced in

    regions where the dose distribution is to be calculated.

    3.5 Material composition

    All materials present in a simulation must be specified by the user. In most

    programs, the materials are specified in terms of their elemental composition and

    density (see chapter 8). Sometimes additional information is required to enhance the

    accuracy of modelling.

    3.6 Source definition

    The tracking of particles starts at a position (or range of positions) where the

    energy and angular distributions of the particles are known with some confidence.

    For instance, in a linac the energy and angular distributions of electrons hitting the

    target are fairly well known. Accelerator modelling is described in more detail in

    chapter 9.

  • 13

    3.7 Scoring

    To extract the absorbed dose distribution from the particle transport simulation,

    one has to define a so-called tally or scoring function. More details on this topic are

    provided in chapter 10.

    3.8 Variance reduction and approximations

    To increase the efficiency of Monte Carlo calculations, approximations and

    variance reduction techniques have been introduced. Examples of approximations

    are the already mentioned condensed history technique for electron transport, and

    the use of cut-off energies. Variance reduction techniques are statistical methods that

    enhance the efficiency of a calculation. Theoretically, these techniques result in

    identical expectation values as without variance reduction, whilst the calculation

    speed is increased. In practice, however, care should be taken and each of these

    techniques should be benchmarked. More details are given in chapter 11.

  • 14

    4 A brief history

    The technique of random sampling to solve mathematical problems is quite old.

    One of the earliest documentations is by Compte de Buffon in 1770. In the early

    nineteen-thirties, using a mechanical adding machine, Fermi already applied

    statistical sampling techniques for radiation transport calculations related to neutron

    diffusion (Metropolis 1987, Wood 1986). The statistical techniques were, however,

    considered impractical as they were time-consuming and tedious. During the second

    world war Mauchly and colleagues developed the first electronic digital computer

    named ENIAC, Electronic Numerical Integrator And Computer, containing around

    18.000 double triode vacuum tubes in a system with half a million solder joints

    (Cooper 1989). Development of the ENIAC was inspired by the labor- and time-

    intensive ballistic computations for generation of firing-tables. The system was

    realised in late 1946, and in 1947 it was moved to its permanent home at the

    Ballistics Research Laboratory in Maryland, USA. Very soon it was realised that the

    ENIAC offered new opportunities for statistical sampling techniques. The first tests

    were on a variety of problems in neutron transport. One of the collaborators, N.

    Metropolis, named the mathematical method Monte Carlo, after the city with its

    famous casinos (Metropolis 1987, Cooper 1989).

    As computers gained speed and memory, the Monte Carlo codes became more

    sophisticated. The first version was written in machine code, but by the early 1960s

    programming languages such as FORTRAN (FORmula TRANslation released in

    1957 by IBM -International Business Machines- and standardised in 1966, 1977 and

    1990) got into use. The fast developments in computer hardware and software and in

    statistics were of great influence on the application of Monte Carlo techniques. These

    Monte Carlo methods on the other hand helped to improve the hard- and software,

    and became one of the most important tools of the statisticians.

    At first, the development of dedicated coupled photon electron transport codes

    for each specific problem required a lot of effort. Today, this is no longer necessary

    due to the availability of general purpose codes, like ETRAN, ITS, MCNP, EGS,

    GEANT, and PENELOPE. Most Monte Carlo systems dedicated to radiotherapy are

    (partially) based on these codes. Therefore, a short history of the most important

  • 15

    general purpose codes is given in the following section. The introduction of Monte

    Carlo into radiotherapy treatment planning is discussed in detail in section 14.1.

    4.1 General purpose codes

    The ETRAN (Electron TRANsport) code, developed and maintained at the

    National Institute of Standards and Technology (NIST), Gaithersburg, Maryland,

    USA, contains the basic algorithms for simulating the tracks of electrons and photons

    travelling through matter (Seltzer 1988). The code was originally developed as a tool

    for solving electron transport problems involving energies up to a few MeV. Later, the

    production and propagation of secondary bremsstrahlung was added, to extend the

    calculation to higher energies. The methods used to generate electron trajectories go

    back to a paper of Berger (1963), describing the sampling from multiple-scattering

    distributions. In the early 1970's, at Sandia National Laboratories, the ETRAN code

    was made more user friendly, especially regarding the specification of the problem

    geometry, and extensions were made to lower energies by including more elaborate

    ionization and relaxation models. The combined software was designated the

    Integrated TIGER Series (ITS) system (Halbleib et al 1988). The Los Alamos

    National Laboratory (LANL) integrated the electron transport algorithms of ITS 3.0

    into their MCNP3 (Monte Carlo Neutron Photon) code, yielding the MCNP4 (Monte

    Carlo N-Particle) system, which was first released in 1990 (Briesmeister 2000).

    Based on this code, a different group at LANL developed MCNPX, which can be

    used to simulate many additional types of particle (Waters 2002).

    During the early 1960's, Nagel wrote his Ph.D. thesis at the Rheinischen

    Friedrich-Wilhelms-Universitt in Bonn on electron-photon Monte Carlo. The in-house

    developed Fortran code was a very practical (freeware) tool for experimental

    physicists during the mid 1960's. Electrons and positrons could be simulated from 1

    GeV down to 1.5 MeV, and photons were followed down to 0.25 MeV. The code was

    limited in geometry handling. From 1972 to 1978, Ford and Nelson from Stanford

    Linear Accelerator Center (SLAC) collaborated to revamp Nagels program and make

    it more user friendly. In addition, special attention was given to allow for easy future

    enhancements. The resulting EGS3 code (Electron Gamma Shower) was introduced

  • 16

    in 1978. Nelson (SLAC) and Hirayama (National Laboratory for High Energy Physics,

    KEK) extended the flexibility of EGS in general, and for high energy accelerators in

    particular. Rogers and colleagues (National Research Council of Canada, NRC)

    extended the code to low energies. These efforts were pooled together in 1985, and

    EGS4 was introduced (Nelson et al 1985). In 1990, PRESTA (Parameter Reduced

    Electron Stepping Algorithm) was introduced in EGS4 (Bielajew and Rogers 1987). In

    2000, Kawrakow and Rogers released the EGSnrc code as the successor to EGS4,

    with further improvements in the modelling of electron transport (Kawrakow and

    Rogers 2000).

    PENELOPE (PENetration and Energy LOss of Positrons and Electrons) was

    developed by Universitat de Barcelona and Institut de Tcniques Energtiques,

    Universitat Politcnica de Catalunya in Barcelona, Spain, and Universidad Nacional

    de Cordoba, Argentina (Salvat et al 2003). It was first released in 1996. PENELOPE

    performs Monte Carlo simulation of electron-photon showers in arbitrary materials.

    Initially, it was devised to simulate the penetration and energy loss of positrons and

    electrons in matter; photons were introduced later. Large efforts were made to make

    the simulation of electron transport as accurate as possible, especially in the low

    energy region.

    The first version of GEANT (GEometry ANd Tracking) was written in 1974 as a

    bare framework, which initially emphasised tracking of a few particles per event

    through relatively simple detectors. The code was developed as a simulation tool for

    high energy physics experiments. From 1993 to 1998, the FORTRAN based

    GEANT3 simulation program was entirely redesigned as an object-oriented program

    written in C++, designated GEANT4 (Agostinelli et al 2003). This code is a

    collaboration of many international research groups under supervision of CERN

    (Conseil Europen pour la Recherche Nuclaire / European Organization for Nuclear

    Research). It is a very versatile code, useful for many different types of particles over

    a wide energy range and capable of handling complex geometries. GEANT4,

    includes a low-energy electromagnetic physics package, which makes it useful for

    radiotherapy applications. Recently, an implementation of the PENELOPE

    electromagnetic physics has also been added to the code.

  • 17

    5 General purpose Monte Carlo codes in radiotherapy

    At present, four general purpose Monte Carlo systems are in use for

    radiotherapy dose calculation. These systems are EGS (Nelson et al 1985,

    Kawrakow and Rogers 2000)), MCNP (Briesmeister 2000, Waters 2002),

    PENELOPE (Salvat et al 2003), and GEANT (Agostinelli et al 2003).

    EGS and PENELOPE simulate the coupled transport of photons and electrons

    (and positrons), while other particles such as neutrons or protons are not taken into

    account. This has the advantage that during the development of these codes all

    attention has been focused on the particles of interest for radiotherapy dose planning.

    On the other hand, in high energy photon beams (18 MV and higher) the production

    of neutrons and protons in the accelerator head may impact (the biological effect of)

    the physical dose distribution in the patient, especially in bone where even alpha

    particles have a non-negligible contribution (Chibani and Ma 2003). These particles

    can be taken into account in MCNP and GEANT. The latter codes were not

    developed specifically for low-energy (radiotherapy) dosimetry, but large efforts have

    recently been made to provide reliable low-energy extensions of these systems.

    In the next paragraphs, the four systems are described in more detail, focusing

    on the mutual differences. In general, it can be said that modelling of photon

    transport is quite similar in all four systems in the energy range of radiotherapy

    applications, although different cross section data are used. The main differences

    occur in the electron transport, which can be dealt with in several ways, having a

    large impact on the speed and accuracy of the systems. In the paragraphs below

    only a short introduction is given. For more details, the reader is referred to the

    corresponding references. An interesting overview has been given by Verhaegen and

    Seuntjens (2003).

    5.1 EGS

    In the past decade, much attention has been paid to the electron transport in

    EGS (Electron-Gamma Shower). In 1990, PRESTA (Parameter Reduced Electron

    Stepping Algorithm) was introduced in EGS4 (Bielajew and Rogers 1987), and in

  • 18

    2000 the EGSnrc code was released by Kawrakow and Rogers as the successor to

    EGS4. In EGS4 (Nelson et al 1985), the Molire (1948) multiple scattering theory is

    used, which is only valid for small scattering angles. In EGSnrc (Kawrakow and

    Rogers 2000, Kawrakow 2000a), an improved multiple scattering theory based on

    screened Rutherford elastic scattering is used instead. Furthermore, this code uses

    PRESTAII (Bielajew and Kawrakow 1997). The main improvement of PRESTAII

    compared to PRESTA is the introduction of a single scattering model of electron

    transport, making it possible to reduce the electron step length to very small values

    near material boundaries. These improvements are expected to improve the

    calculation accuracy of angular deflections for electrons, eliminate restriction on the

    maximum and minimum electron path length in EGS4/PRESTA-I imposed by the

    Molire theory, and provide an exact boundary-crossing algorithm by using single

    elastic collisions of electrons.

    From the benchmarks applied to EGSnrc (Kawrakow 2000b, Verhaegen 2002),

    it can be concluded that this code is very accurate even in the vicinity of interfaces

    between materials with high and low atomic numbers (Z). However, for MCTP

    applications EGS4 (PRESTA) seems good enough and is faster than EGSnrc. A

    disadvantage of EGS4 and EGSnrc is that users need to program their code in a

    macro Fortran code called Mortran. Obviously, only the geometry, source input, and

    tallying need to be programmed. In a pre-compilation step, the user code is

    connected to the EGS core.

    Two user codes, designated BEAM and DOSXYZ (Rogers et al. 1995, Rogers

    et al 2002), are available for applications in MCTP. BEAM is an EGS user code

    specifically developed for the modelling of a linear accelerator. All components of the

    accelerator (target, primary collimator, flattening filter, monitor, jaws, MLC, etc.) are

    pre-programmed in so-called component modules. The user can build an accelerator

    by simply summing the required components. An input file must be generated in

    which the dimensions, materials and transport parameters of the individual

    components must be defined. No programming efforts are required. With BEAM it is

    possible to determine so-called phase-space files in a plane at the exit of the linear

    accelerator. These files contain all necessary parameters (direction, location, energy,

    charge, etc.) of particles passing through the plane. Such files can then be used as

  • 19

    input for dose calculations in phantoms or patients using the other pre-programmed

    user code, designated DOSXYZ. In this code CT data can be imported and translated

    to voxels with a certain material and density. Systems as MCDOSE, Peregrine,

    XVMC and DPM (section 13) are totally or partially based on BEAM and DOSXYZ.

    5.2 MCNP

    MCNP is a general-purpose, continuous-energy, generalised-geometry, time-

    dependent, coupled neutron/photon/electron Monte Carlo transport code. Two

    versions of the MCNP (Monte Carlo N-Particle) code, developed by different groups,

    currently exist. MCNP4C (Briesmeister 2000), is able to simulate the (coupled)

    transport of neutrons, photons and electrons, whereas MCNPX (Waters 2002) can

    simulate a variety of other particles as well. The photon and electron physics in the

    present version of MCNPX (version 2.5) are identical to those in MCNP4C. Hence, in

    the following we will denote both codes as MCNP. It is noted that the successor of

    MCNP4C, MCNP5 (Brown 2003), has been released, but is not yet available outside

    the USA.

    The electron transport algorithms in MCNP are claimed to be equal to those in

    the ITS 3.0 system (Halbleib et al 1988), which in turn were derived from ETRAN

    (Seltzer 1988). The Goudsmit-Saunderson multiple scattering theory is used, while

    the sampling of energy loss is based on the Landau straggling theory. Several

    investigators have shown though that care should be taken with the electron

    transport (Jeraj et al 1999, Schaart et al 2002, Reynaert et al 2002). A systematic

    error is present in the default MCNP electron energy indexing algorithm. However,

    the user can choose to use the ITS electron energy indexing algorithm instead, which

    leads to correct results. An additional problem exists with MCNP4C when the

    geometry contains many boundaries, e.g. in the case of a voxelised phantom.

    MCNP4C requires the voxels in such a phantom to be modelled as separate material

    regions, even if they exist of the same material. It has been shown that in such cases

    the cumulative effect of many small boundary crossing artefacts may lead to

    significant errors in the calculated dose distribution (Schaart et al 2002, Reynaert et

    al 2002).

  • 20

    In contrast to EGS and GEANT4, MCNP does not require any programming by

    the user. Instead, the user only needs to provide an ASCII input file specifying the

    problem geometry (using a variety of available surface types and/or macrobodies

    such as spheres, boxes and cylinders), the source(s) (energy and angular spectra,

    etc.), the tallies (e.g. energy deposition or track length), and (optionally) the use of

    one or more of the many available variance reduction techniques. The simulation

    results are provided in ASCII output files. Graphical user interfaces, such as VISED

    (2004) are available to generate input files and to visualise the output data.

    5.3 PENELOPE

    PENELOPE (PENetration and Energy LOss of Positrons and Electrons) has

    been introduced recently (Sempau et al 1997, Salvat et al 2003). The code simulates

    the coupled transport of electrons, positrons and photons with energies between a

    few hundred eV and 1 GeV. It is capable of handling complex geometries and static

    electromagnetic fields. Large efforts were made to make the simulation of electron

    transport as accurate as possible. Ideas introduced in PENELOPE have been

    implemented in EGSnrc and vice versa. So it can be expected that these codes will

    provide rather similar results. In PENELOPE a mixed scheme of single and multiple

    scattering is used, comparable to EGSnrc. The multiple scattering algorithms are

    based on the Goudsmit-Saunderson theory. In the PENELOPE implementation of

    multiple scattering, the angular deflection and the lateral displacement for each

    electron step are accounted for using the so-called random hinge method, which is a

    simple and fast method for obtaining an accurate geometric representation of the

    electron track. The user has to program the application in Fortran, although several

    user codes are available in the system. Benchmarks of PENELOPE against other

    codes and experiments have recently been published by Sempau et al (2001),

    Sempau et al (2003) and Ye et al (2004). These studies generally show good

    agreement with EGS and experiments. The applicability for linac modelling has been

    illustrated in Sempau et al (2003).

  • 21

    5.4 GEANT

    GEANT (GEometry ANd Tracking) was originally developed for high-energy

    physics. It can be used for the simulation of many types of particle over a wide

    energy range. The current version, GEANT4, includes a low-energy electromagnetic

    physics package, which makes it useful for radiotherapy applications (Agostinelli et al

    2003). Recently, an implementation of the PENELOPE electromagnetic physics has

    also been added to the code. The code can handle complex geometries,

    electromagnetic fields, (electronic) detector response, and allows for time-dependent

    (4D) modelling of e.g. decaying particles and/or moving objects. A variety of

    visualization tools is provided, as well as connectivity to data-analysis software and

    computer-aided design (CAD) programs (for geometry input). The user must provide

    a set of C++ objects that are built upon the Monte Carlo core of the program in an

    object-oriented approach.

    Recently, GEANT4 has found use in a variety of medical physics applications

    (Barca et al 2003, Archambault et al 2004). Some benchmarks of GEANT4 electron

    and photon transport against other Monte Carlo codes and measurements have been

    published by Carrier et al (2004) and Rodriques et al (2004). These studies showed

    good agreement for photons. Carrier et al reported fair agreement for electrons,

    although some non-negligible differences with e.g. EGSnrc (4% for a 10 MeV parallel

    beam) were found (see also Torres et al 2004). Recently Poon and Verhaegen

    (2005) extensively benchmarked GEANT4 against EGSnrc for radiotherapy

    applications. In this paper, a very nice overview of the photon and electron transport

    physics modelled in the GEANT code is presented for the 3 different electromagnetic

    physics models (standard, low-energy, Penelope). For photon beams depth dose

    curves are in good agreement except in the buildup zone. For electron beams

    differences are more important. It is also illustrated that results depend highly on

    transport parameters as e.g. the electron step size. This is even more clearly

    demonstrated in the paper of Poon et al (2005), where a more fundamental study of

    the electron transport in GEANT4 is performed. Accurate results can be obtained

    after careful selection of transport parameters. In that case the code is an order of

    magnitude slower than e.g. EGSnrc. As new releases of GEANT4 are continuously

  • 22

    improved with respect to the code, it can be expected that the role of GEANT4 in

    medical physics may become more important in the near future.

    In this context it is interesting to note that the OpenGATE collaboration has

    recently released the first version of GATE, a modular, scripted, GEANT4-based

    Monte Carlo code which, in contrast with GEANT4 itself, does not require the user to

    be familiar with C++ (Jan et al 2004). Although this code was primarily developed for

    nuclear medicine applications (modelling of PET and SPECT scanners), extensions

    into other domains such as radiotherapy are currently being developed.

  • 23

    6 Rationale for Monte Carlo treatment planning

    6.1 Requirements on uncertainty in Treatment Planning

    An interesting discussion on uncertainty in treatment planning is provided in

    AAPM report No 85 of the AAPM Task Group 65 (Papanikolaou et al 2004). As

    stated in this report, due to the steep slope of the TCP-and NTCP-dose relationships,

    a dose error of 5 % might lead to a TCP change of 10% to 20%, and to even larger

    NTCP changes (see also Fraass et al 2003). Clinical effects are already noticeable

    for dose errors of 7 % (Papanikolaou et al 2004). Therefore accurate dose

    information is required.

    Between the dose prescription to a tumour and the actual dose delivery a large

    number of steps are involved. During each step, uncertainties are introduced,

    accumulating to an overall uncertainty for the full process of dose delivery. An

    overview of the various components of uncertainty is given in Table 1 of AAPM

    Report 85. An overall uncertainty of 4.3 % (1) is obtained, which is in

    correspondence with the more familiar 5 % (1) obtained in previous work (Mijnheer

    et al. 1987, ICRU 1976).

    Improving the quality of the dose engine, i.e. reducing the uncertainty in the

    dose calculation, will reduce the overall uncertainty in the delivered dose. It should be

    noted that the use of an extremely accurate dose engine will not automatically lead to

    very low uncertainties in clinical dose delivery as several other factors contribute

    significantly to the overall uncertainty. However, in AAPM report 85 it is claimed that

    the overall uncertainty in the delivered dose will decrease to 2.5 % (1), leading to a

    situation where the accuracy of the dose engine plays an important role. At present, it

    is generally believed that the dose calculation should be accurate to within 2% - 3%

    (1) (Fraass et al 2003).

  • 24

    6.2 Why Monte Carlo Treatment Planning

    Monte Carlo dose calculation engines have the potential to meet, or even

    perform better than, the 3 % (1) uncertainty requirement, regardless of beam

    geometry and patient composition. As for any type of dose engine, however, the

    uncertainty for a Monte Carlo dose engine will never be zero due to, for example:

    imperfect matching of the Monte Carlo beam to the actual accelerator beam,

    uncertainties in the cross section libraries,

    the standard deviation due to the limited number of histories simulated,

    uncertainties in the conversion of CT data to material composition and density.

    The quality of beam matching is very difficult to estimate, but in general it should

    be possible to achieve this within 1 % (1) or better (Verhaegen and Seuntjens 2003

    and Ma, Jiang 1999). Most authors assume that the uncertainty in cross section

    libraries is small enough to be negligible (Fraass et al 2003). The statistical

    uncertainty depends on the number of histories. The uncertainty associated with

    tissue characterization is difficult to quantify. Instead of using water with different

    densities for all tissue types, the real tissue composition must be estimated for the

    calculation of cross sections.

    Taking all of the above-mentioned uncertainties into account, Monte Carlo

    treatment planning is expected to be able to offer an uncertainty in dose calculation

    well within 3 % (1) required for accurate radiotherapy. Other advantages are given

    by Fraass et al (2003). One advantage over conventional dose engines is that the

    uncertainties are independent of the treatment setup. Furthermore, the Monte Carlo

    method could lead to an increase in confidence in the obtained dose distributions

    (see also Cygler et al 2005). This could lead to the delivery of a higher tumour dose

    to avoid recurrence, while having faith in the reported dose to critical organs.

    An interesting discussion is provided in a point/counterpoint discussion between

    Mohan and Antolak (2001). Arguments against MCTP raised by Antolak include: the

    influence of (statistical) noise, the influence of approximations and variance reduction

    techniques introduced to limit the calculation time and the limited spatial resolution

    (voxel size) often used, again to speed up the calculations. These arguments are

    considered of minor importance by Mohan: approximations and variance reduction

  • 25

    techniques are illustrated to introduce no bias, the effect of statistical noise is very

    limited and resolutions up to 2 or 3 mm can be reached within a few minutes of

    calculation time. It is clear, however, that the added value of MCTP compared to

    superposition/convolution algorithms should be illustrated by examples. In the

    following two paragraphs a literature study of phantom studies and comparisons for

    clinical cases is provided.

    6.3 Phantom experiments

    In the vicinity of low density volumes (lung) and air cavities, Monte Carlo dose

    calculations have been reported to be more accurate than conventional techniques

    (Mohan et al 1997, Solberg et al 1998, Ma et al 1999, Keall et al 2000, Martens et al

    2002, Heath et al 2004, Paelinck et al 2005). Mohan et al (1997) stated that

    conventional methods (including superposition/convolution techniques) will give rise

    to deviations ranging from 5 % to 10 % in the presence of tissue heterogeneities. The

    results of Ma et al (1999) illustrate that MCTP is certainly interesting for electron

    beams, as e.g. the FOCUS conventional dose calculation algorithm (pencil beam

    algorithm) leads to large deviations (up to 15 %) and isodose line shifts of more than

    1 cm (see figure 6.1).

  • 26

    Figure 6.1: Isodose line shift between results obtained with the FOCUS pencil

    beam algorithm (a) and Monte Carlo calculations (b) (reproduced with kind

    permission of AAPM from Ma et al (1999)).

    For photon IMRT applications, an added value of the MC method can be found

    in head-and-neck treatment and treatment of lung cancer, because of the presence

    of tissue inhomogeneities resulting in loss of electronic equilibrium. For IMRT the

    best available non-Monte Carlo dose calculation engines are based on the

  • 27

    superposition/convolution method (Boyer and Mok 1984, Mackie et al 1985, Ahnesj

    1989, Keall and Hoban 1996, Yu et al 1995). Ma et al (1999) obtained large

    differences between the FOCUS planning system and a Monte Carlo dose engine for

    a phantom containing lung or bone layers, even when the superposition convolution

    method of FOCUS was used. An interesting comparison of two

    superposition/convolution algorithms and the Monte Carlo method for a lung cavity is

    provided by Paelinck et al (2005) (see figure 6.2).

    Figure 6.2: Comparison of two superpostion/convolution algorithms and Monte

    Carlo calculations for a phantom with a lung insert in a 6 MV beam (reproduced with

    kind permission from Paelinck et al (2005)).

    The Helax TMS system (Nucletron, Veenendaal, The Netherlands)

    systematically underestimates the dose in the lung-equivalent cavity by 6 %, while

  • 28

    the Pinnacle algorithm (Philips Medical Systems, Best, the Netherlands)

    overestimates the dose behind the cavity by 4 %. Also in the work of Crammer-

    Sargison et al (2004), significant deviations in lung equivalent material were obtained

    for the CadPlan pencil beam convolution algorithm (Varian Oncology Systems Inc.,

    Palo Alto, CA). Arnfield et al (2000) obtained substantial deviations between

    measurements and superposition/convolution (Pinnacle) in and around lung-

    equivalent material, while the Monte Carlo results are in excellent agreement with the

    measurements. These deviations become more important when simulating a small

    (4x4 cm) high energy photon beam (18MV). Krieger and Sauer (2005) performed a

    comparison between the pencil beam (Helax TMS), superposition/convolution (Helax

    TMS) and Monte Carlo methods for a multi-layer phantom consisting of styrofoam (to

    simulate the low density of lung) and polystyrene layers for regular beams. In

    polystyrene, superposition/convolution and MC were in agreement with the

    measurements while the pencil beam algorithm deviated by 12 %. In styrofoam,

    however, even the superposition/convolution algorithm deviated by more than 8 %

    from measurements and MC results.

    6.4 Comparisons for clinical cases

    In the examples described above, extreme situations were investigated

    consisting of one single beam crossing a large lung/air cavity. It is not straightforward

    to extrapolate these findings to clinical practice. Therefore in this paragraph we will

    focus on examples of realistic clinical calculations. The results are discussed

    chronologically and the focus is on the most recent results as these are obtained with

    the most recent (and thus most accurate) versions of the available conventional dose

    calculation engines.

    Wang et al (1998) developed a patient specific Monte Carlo dose engine that

    was evaluated for conformal lung treatment. The method was approximate as only

    one medium (water) was defined, although density variations where taken into

    account. The dose distributions obtained were compared against a conventional dose

    engine based on the equivalent path length (EPL) method. The Monte Carlo results

    illustrated that 20 % of the planning target volume (PTV) was underdosed, while the

  • 29

    maximum doses in cord and heart (two parameters used in the objective function of

    the treatment planning system) were underestimated by the conventional system by

    more than 25 %. Deviations were attributed to the approximate modelling of lateral

    particle transport in low density regions by the conventional dose calculation engine.

    In a follow-up study (Wang et al 2002), the same PB algorithm and MC code were

    compared for IMRT treatment of five lung patients and four head-and-neck patients.

    For one lung patient, a decrease of 10% in D95 and 6 % in Dmean was obtained, while

    for the other patients the PTV coverage decreased with 2-5%. For one of the head-

    and-neck patients (a patient with recurrence) D95 differed by 9 %. In lung, differences

    in D05 and Dmax of up to 10 % were found. Also in the spinal cord, differences larger

    than 5 % were noticed. For all head-and-neck patients, dose differences in the optical

    chiasm were below 2 %. An interesting conclusion is that larger effects are observed

    for individual fields than for the composite plan.

    Figure 6.3: Comparison between Konrad Pencil beam calculations and EGSnrc

    for head and neck treatments (reproduced with kind permission from Laub et al

    (2000)). DVHs are for the PTV and optical chiasm. Abbreviations used: PB (Pencil

    Beam), Veri (EGS4 Verification calculation), IM/MC (intensity modulated/Monte

    Carlo). IM/MC is the dose distribution obtained from the Monte Carlo inverse planning

    system.

  • 30

    Laub et al (2000) obtained large differences between the KonRad Pencil Beam

    algorithm with 1D inhomogeneity correction (and accounting for lateral electron

    transport) on one hand and EGS4 (Nelson et al 1985) on the other hand for head-

    and-neck treatments (see figure 6.3).

    The Monte Carlo result in the PTV was systematically lower than the PB result,

    although it is not clear whether the MC dose was expressed as dose to water

    (presence of large air cavity with corresponding low stopping powers in the PTV can

    lead to differences in DVH of PTV, see par 10.4 for a more detailed explanation).

    Differences were attributed to the rebuild-up behind the air cavity. According to the

    MC results the dose constraint in the chiasm was violated.

    Francescon et al (2000) compared the superposition/convolution algorithm of

    Pinnacle with the Monte Carlo code BEAM (Rogers et al 1995) for mediastinal and

    breast treatments. Deviations were below 2.5 % and thus within 2 standard

    deviations of the Monte Carlo calculation. Also for single fields and large

    inhomogeneities the differences were negligible. The study was restricted to large

    beams. As stated by Ahnesj (1989) larger deviations are expected for smaller fields.

    Jeraj et al (2002) illustrated that two types of error are introduced when using an

    approximate dose calculation algorithm for inverse treatment planning, namely a

    systematic error due to errors in the dose calculations and a convergence error

    resulting from the fact that the optimised beam settings obtained by the approximate

    dose engine will differ from those obtained with an accurate dose calculation

    algorithm. In this study, results obtained by Monte Carlo, superposition/convolution

    and pencil beam methods were compared. Systematic errors were below 1% of Dmax

    in the tumour and slightly larger outside the PTV for the superposition/convolution

    method and around 5% for the pencil beam algorithm. The authors concluded that

    pencil beam algorithms should be replaced by superposition/convolution or Monte

    Carlo algorithms.

    Leal et al (2003) compared the Plato PB algorithm (Nucletron, Veenendaal, The

    Netherlands) with the Monte Carlo program BEAM for different clinical cases. As

    illustrated in figure 6.4, significant differences were obtained when comparing the

    DVHs in the bladder and the rectum.

  • 31

    Figure 6.4: Plato PB algorithm (TPS) versus Monte Carlo (MC) for a prostate

    treatment (reproduced with kind permission from Leal et al (2003)).

    As recently stated by Chetty et al (2005): when comparing Monte Carlo results

    with conventional dose calculation engines, it would be interesting to distinguish

    between effects related to differences in the beam model and effects related to the

    particle transport within the patient geometry. Therefore Chetty et al used two

  • 32

    versions of an equivalent path length algorithm, namely a version with an

    approximate beam model, and one with an accurate beam model that provides

    excellent agreement when comparing calculational results with measurements in a

    homogeneous phantom.

    Figure 6.5: Comparison of equivalent path length (EPL) algorithm with Monte

    Carlo calculations (DPM) illustrating the importance of accurate tuning for a

    homogeneous phantom (above) and a heterogeneous phantom (below). The results

    depicted with best fit are obtained with the accurate beam model. (reproduced with

    kind permission from Chetty et al (2005)).

  • 33

    These two models were compared with the DPM (see section 14.2 for a

    description of DPM) Monte Carlo dose engine for a homogeneous phantom, a

    heterogeneous thorax phantom and a lung patient plan (see figure 6.5).

    The importance of an accurate beam model was illustrated by the fact that the

    EPL algorithm using the accurate beam model (best-fit results) gave rise to a much

    better agreement with the Monte Carlo results for 6 MV in a homogeneous phantom.

    For the lung phantom though, the disagreement between the mean lung dose of the

    best-fit results and the MC method was 30 % for 15 MV. This illustrates (as stated by

    the authors) that especially at high energy (15 MV) the inhomogeneity effects

    (transport of secondary electrons in low density regions) may be more significant

    than beam model approximations.

    Figure 6.6: Comparison of Corvus pencil beam algorithm with MCSIM Monte

    Carlo calculations for the prostate (reproduced with kind permission from Yang et al

    (2005)).

    Yang et al (2005) compared the Corvus finite-size PB algorithm (with and

    without inhomogeneity corrections) with Monte Carlo calculations (MCSIM, see

  • 34

    section 14.3) for 25 coplanar and 5 non-coplanar IMRT plans for the prostate (see

    figure 6.6).

    For the coplanar plans the agreement between MCSIM and Corvus was within 3

    %. For the non-coplanar plans, differences up to 7 % in Dmean and above 8 % in D98

    were obtained in the PTV. Another conclusion was that it was necessary to apply the

    EPL heterogeneity corrections in Corvus.

    Boudreau et al (2005) compared Corvus (with and without EPL correction) with

    the Peregrine Monte Carlo method (see section 14.5) for IMRT head and neck

    treatment planning (see figure 6.7 and table 6.1).

    Table 6.1: Summary of ratios between Corvus and Peregrine results obtained

    by Boudreau et al (2005) (reproduced with kind permission of Boudreau et al (2005)).

  • 35

    Figure 6.7: Comparison of Corvus system with Peregrine Monte Carlo

    calculations (reproduced with kind permission from Boudreau et al (2005)).

    For the brainstem, Peregrine delivered on average a 6% higher Dmean, so for

    individual patients even larger differences were obtained. The Peregrine system was

    extensively benchmarked against measurements (Heath et al 2004).

    Reynaert et al (2005) presented a comparison between two Monte Carlo dose

    calculation engines (Peregrine and MCDE) and the Helax TMS

    superposition/convolution algorithm for a head-and-neck patient (see figure 6.8).

  • 36

    (a)

    (c)

    Figure 6.8: Comparison between the Monte Carlo dose calculation engines

    Peregrine and MCDE. In part (a) the MCDE doses (obtained with the MLCE model

    for the Elekta MLC) were systematically multiplied by 1.07, illustrating a dose

    difference of 7 % in the optical chiasm. In part (b) (lateral profiles of 2x40 and 40x2

    beam segments) the cause of the discrepancies is demonstrated (problem with MLC

    model). (reproduced with kind permission from Reynaert et al 2005).

    (b)

  • 37

    In this work it was demonstrated that Peregrine provided systematic errors in

    the DVHs in the optical chiasma, due to a systematic error in the leaf projection. The

    superposition/convolution results are in acceptable agreement with MCDE. Only one

    patient was studied.

    Figure 6.9: Comparison of PB algorithm, superposition/convolution (Helax TMS)

    and BEAMnrc/DOSXYZnrc Monte Carlo calculations for a head and neck patient.

    (reproduced with kind permission from Seco et al (2005))

  • 38

    Seco et al (2005) performed a comparison between a PB, a

    superposition/convolution (Helax TMS) and a MC system for a head and neck patient

    (see figure 6.9).

    Large differences were obtained in the PTV but, as stated by the author, this

    was largely caused by the fact that the MC dose in the air cavities was expressed as

    dose to medium (see section 10.4). In the critical structures the Monte Carlo DVHs

    differed significantly from the PB and superposition/convolution results.

    The most direct way to determine the added value of MCTP is to try to link

    observed differences in dose maps to clinical outcome. This can be done by e.g.

    comparing post-treatment CT scans with (1) possible recurrence within the PTV with

    regions of underdosage (as predicted by the MC results) and (2) possible side effects

    in critical tissues in regions of overdosage. Data on this topic is still missing. An

    interesting paper on this topic was published by De Jaeger et al (2003). Lung cancer

    patients, originally planned with an EPL algorithm, were retrospectively recalculated

    with a superposition/convolution algorithm, illustrating large differences in the mean

    lung dose (up to 20 %). The Lyman model was used to illustrate that these dose

    differences can lead to complications in lung tissue.

    6.5 Conclusions

    Based on single beam phantom experiments it can be concluded that well-

    benchmarked MC dose engines clearly outperform both PB and

    superposition/convolution algorithms regarding dosimetric precision. Also for realistic

    clinical plans the MC codes are superior to PB calculations. More studies are needed

    to investigate to what extent the replacement of superposition/convolution algorithms

    by MC may result in a benefit for clinical plans.

    Most published MC results for photon beams were obtained by MC experts with

    well benchmarked research systems. The influence of the approximations and

  • 39

    variance reduction methods introduced in commercial MCTP systems on the

    uncertainty are not yet clear. Moreover, even an MCTP system without

    approximations or variance reduction methods can contain systematic errors.

    Consequently, every individual MCTP system must be benchmarked before clinical

    use.

    An important conclusion is that Monte Carlo dose calculation engines, when

    carefully validated against measurements, provide an additional benchmarking tool

    for treatment planning, in situations where measurements are difficult or even

    impossible (Mohan 1988).

  • 40

    Part II: Fundamentals of Monte Carlo

  • 41

    7 Modelling of particle transport

    The following discussion will be restricted to coupled photon-electron transport

    as this is the focus of the present report. In a recent paper, Chibani and Ma (2003)

    investigated the influence of photonuclear reactions in the linac head for high-energy

    photon beams (18 MV and higher). The effects of neutrons, protons and alphas on

    dose (taking into account the RBE of the particles) are below 0.7 %. Therefore it is

    unlikely that these particles will ever be taken into account in a MCTP system for

    photon and electron beams.

    7.1 Photon transport

    In general, the types of photon interaction taken into account in a Monte Carlo

    treatment planning code are the photoelectric effect, Compton scattering, Raleigh

    scattering and pair production.

    In the case of photoelectric absorption, the photon interacts with a (tightly

    bound) atomic electron. In this process, which is dominant at low photon energies,

    the photon disappears and all of its energy is transferred to the electron, which is

    ejected from the atom with a kinetic energy equal to the difference between the initial

    photon energy and the electrons binding energy. As a result of this process, one of

    the atomic shells is left with a vacancy which is promptly filled by a less tightly bound

    electron, resulting in the emission of a fluorescence X-ray or one or more Auger

    electrons. In a detailed Monte Carlo simulation, all of the secondary particles (photo-

    electrons, X-rays and Auger electrons) may be transported. However, in cases where

    this does not significantly influence the end result, computing time may be saved by

    switching off the transport of one or more of these types of secondary particles.

    In case of Compton scattering, a photon interacts with a free (i.e. unbound)

    electron. If the photon energy is high with respect to the binding energy of an electron

    in its atom, this electron can be considered free for this purpose. Part of the photon

    energy is transferred to the electron. The scattered photon and the electron emerge

    from the interaction at angles relative to the direction of the initial photon that are

    related to the particle energies because of the conservation of energy and

  • 42

    momentum. Compton scattering is an important process for the energies of interest in

    radiotherapy, especially in low-Z materials.

    In Raleigh scattering, essentially no energy is exchanged; only the direction of

    the photon is changed, usually by a small angle. Raleigh scattering is also called

    coherent scattering since the photon scatters elastically off an entire atom, where all

    electrons behave coherently. The importance of Raleigh scattering is relatively small,

    but not always negligible.

    In the case of pair production, the photon disappears and an electron-positron

    pair is created. This process is only possible if the photon energy is higher than twice

    the electron rest mass (2 511 keV), and dominates at high energies in dense

    materials. The positron created in the interaction will annihilate with an electron when

    it comes to rest, resulting in the emission of two 511 keV annihilation photons.

    In Section 1, an example is given on how the transport of photons is simulated.

    It is explained that, for each photon emitted by the source, the distance to the first

    interaction is sampled, based on the probability exp(- l) that the photon will not

    interact over a distance l.

    The photon is then transported to the location of the first interaction.

    Subsequently, the type of interaction to be simulated is sampled, based on the partial

    cross sections for the different interactions contained in the interaction data tables

    (see section 7.3). The selected type of interaction is then simulated. Here, use is

    made of the well-known theories describing the kinematics of the various types of

    photon interaction, see e.g. Attix (1986). In case of, for example, a Compton

    interaction, the energy and direction for the scattered photon are sampled. If electron

    transport is taken into account, the energy and direction of the electron participating

    in the interaction are also calculated. This particle is put on the stack for later

    transport. Then, the distance to the next interaction is sampled for the Compton

    scattered photon, and the process is repeated until the photon is absorbed and all

    secondary particles have been transported.

    If a photon is transported through a phantom consisting of multiple materials, it

    is possible that the sampled distance to the next interaction exceeds the distance to

    the nearest material boundary. In such cases, the photon is first transported to the

    boundary. Then, the distance to the next interaction is sampled using the cross

  • 43

    sections of the material into which the photon is entering. The photon track is then

    continued into the new material region (without changing the direction of flight).

    In some calculations, the transport of the fast electrons created by photons can

    be ignored since they transport energy over negligibly small distances and/or

    charged-particle equilibrium exists. Since electron transport tends to consume a lot of

    computation time in Monte Carlo simulations, it may be attractive to switch off

    electron transport in such cases. However, the fast electrons may in turn produce

    (bremsstrahlung or X-ray) photons, which may have a significant effect on the end

    result. Therefore, some Monte Carlo codes offer the possibility to generate such

    secondary photons even if electron transport is turned off. The algorithms used for

    this purpose rely on certain assumptions (e.g., in the tick-target approximation it is

    assumed that each secondary electron is completely absorbed in the same material

    in which the corresponding photon interaction has taken place) and therefore need to

    be used with some caution.

    The physics of photon transport is implemented very similarly in most modern

    Monte Carlo codes. Small details can nevertheless be different, e.g. the handling of

    the Compton effect regarding the binding of the atomic electron. A condensed

    overview of these differences can be found in Verhaegen and Seuntjens (2003).

    7.2 Electron transport

    The physical processes to be modelled when simulating the transport of

    electrons through matter are elastic scattering by (screened) atomic nuclei, inelastic

    collisions with atomic electrons causing either excitation or ionisation,

    Bremsstrahlung production, and the emission of X-rays and Auger electrons following

    electron-impact ionisation. Nuclear processes (which only occur at high electron

    energies) are often neglected. Positrons are sometimes simply modelled as electrons

    with the addition that annihilation photons are created when the particle comes to

    rest. More elaborate models use separate positron cross-section tables and include

    rare positron decay processes such as in-flight annihilation and three-photon

    annihilation.

    An important difference between modelling of electrons and photons lies in the

    fact that photons undergo a relatively small number of discrete interactions per

  • 44

    particle track, whereas electrons undergo a very large number of Coulomb

    interactions with the electrons and atomic nuclei in the material. It is computationally

    very expensive to simulate each of these individual Coulomb interactions, and

    therefore this is not normally done in general-purpose codes or dose engines for

    treatment planning.

    Instead, a so-called condensed-history approach is usually applied (Berger

    1963). In such a model, each electron track is subdivided into a series of short track

    segments, usually called steps. Instead of modelling the individual elastic and

    inelastic collisions along each step, the resulting (cumulative) energy loss and

    angular deflection are sampled once per step only.

    The sampling of angular deflection may be based on a so-called multiple-

    scattering formalism. One example is the implementation, in EGS4, of the theory by

    Molire (1948). The Molire distribution is a universal function of a scaled angular

    variable, which makes it relatively easy to sample the angular deflection for arbitrary

    step lengths during a run. A disadvantage of this theory is that it is based on a small-

    angle approximation, so large-angle deflections are modelled less accurately.

    Another multiple-scattering theory, the Goudsmit-Saunderson (1940) formalism, is

    valid for all scattering angles. However, sampling the angular deflection for arbitrary

    step lengths during a run is less straightforward, so codes based on this theory (such

    as ETRAN, ITS and MCNP) usually sample the deflection angle from stored multiple-

    scattering distributions that have been calculated for a pre-selected set of path

    lengths during the initiation phase of the run (Berger and Wang 1988).

    The sampling of electron energy loss may be done in different ways. A

    distinction is commonly made between so-called class I and class II algorithms

    (Berger 1963, Rogers and Bielajew 1988), see Figure 7.1.

    In a class I code the primary electron is not directly influenced by the generation

    of a secondary electron. Instead, energy straggling (i.e., the fluctuation in electron

    energy due to differences in the energy lost by different electrons of equal initial

    energy traversing the same path length) due to the creation of secondary electrons is

    taken into account explicitly in the algorithm used to sample the energy loss for each

    electron step. Examples of such codes are ETRAN and MCNP, in which the energy

    loss is sampled from the Landau (1944) straggling distribution. An advantage of this

  • 45

    approach is that energy straggling is always modelled accurately, even if a high

    energy threshold for knock-on production is applied. This may greatly speed up a

    simulation if the transport of low-energy secondary electrons is not important. A

    disadvantage of the class I approach is the possibility for negative energy loss events

    in small voxels. Such events may occur if the energy carried out of a voxel by a

    secondary electron created within it is larger than the amount of energy deposited in

    the voxel by the primary (and secondary) electron.

    Figure 7.1 Different ways to perform a sampling of electron energy loss, class I